proteomic analysis of marinobacter hydrocarbonoclasticus

29
HAL Id: hal-01614789 https://hal.archives-ouvertes.fr/hal-01614789 Submitted on 18 Feb 2018 HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers. L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés. Proteomic analysis of Marinobacter hydrocarbonoclasticus SP17 biofilm formation at the alkane-water interface reveals novel proteins and cellular processes involved in hexadecane assimilation Pierre-Joseph Vaysse, Laure Prat, Sophie Mangenot, Stéphane Cruveiller, Philippe Goulas, R. Grimaud To cite this version: Pierre-Joseph Vaysse, Laure Prat, Sophie Mangenot, Stéphane Cruveiller, Philippe Goulas, et al.. Proteomic analysis of Marinobacter hydrocarbonoclasticus SP17 biofilm formation at the alkane-water interface reveals novel proteins and cellular processes involved in hexadecane assimilation. Research in Microbiology, Elsevier, 2009, 160 (10), pp.829-837. 10.1016/j.resmic.2009.09.010. hal-01614789

Upload: others

Post on 16-Mar-2022

4 views

Category:

Documents


0 download

TRANSCRIPT

HAL Id: hal-01614789https://hal.archives-ouvertes.fr/hal-01614789

Submitted on 18 Feb 2018

HAL is a multi-disciplinary open accessarchive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come fromteaching and research institutions in France orabroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, estdestinée au dépôt et à la diffusion de documentsscientifiques de niveau recherche, publiés ou non,émanant des établissements d’enseignement et derecherche français ou étrangers, des laboratoirespublics ou privés.

Proteomic analysis of Marinobacterhydrocarbonoclasticus SP17 biofilm formation at the

alkane-water interface reveals novel proteins and cellularprocesses involved in hexadecane assimilation

Pierre-Joseph Vaysse, Laure Prat, Sophie Mangenot, Stéphane Cruveiller,Philippe Goulas, R. Grimaud

To cite this version:Pierre-Joseph Vaysse, Laure Prat, Sophie Mangenot, Stéphane Cruveiller, Philippe Goulas, et al..Proteomic analysis of Marinobacter hydrocarbonoclasticus SP17 biofilm formation at the alkane-waterinterface reveals novel proteins and cellular processes involved in hexadecane assimilation. Researchin Microbiology, Elsevier, 2009, 160 (10), pp.829-837. �10.1016/j.resmic.2009.09.010�. �hal-01614789�

1

Proteomic analysis of Marinobacter hydrocarbonoclasticus SP17 biofilm formation at the 1

alkane-water interface reveals novel proteins and cellular processes involved in 2

hexadecane assimilation. 3

4

Pierre-Joseph Vayssea, Laure Prata†, Sophie Mangenotb, Stéphane Cruveillerc, Philippe 5

Goulasa, Régis Grimauda* 6

7

a Institut Pluridisciplinaire de Recherche en Environnement et Matériaux, Equipe 8

Environnement et Microbiologie UMR5254 CNRS, IBEAS, Université de Pau et des Pays de 9

l’Adour, BP1155, 64013 Pau cedex, France 10

11

b CEA/DSV/IG/Genoscope, 2 rue Gaston Cremieux, 91057 Evry cedex, France 12

13

c CEA/DSV/IG/Genoscope/LGC, 2 rue Gaston Cremieux, 91057 Evry cedex, France 14

15

† Current address : EPFL, ENAC-ISTE, Laboratory of Environmental Biotechnology, 16

Lausanne, Switzerland 17

18

E-mail addresses of all authors 19

• Pierre-Joseph Vaysse: [email protected] 20

• Laure Prat: [email protected] 21

• Sophie Mangenot: [email protected] 22

• Stéphane Cruveiller: [email protected] 23

• Philippe Goulas: [email protected] 24

• Régis Grimaud: [email protected] * Correspondence and reprints 25

Tel: 33 (0)5 59 40 74 86, Fax: 33 (0)5 59 40 74 94 26

27

2

Abstract 28

Many hydrocarbon degrading bacteria form biofilms at the hydrocarbon-water interface to 29

overcome the low accessibility of these poorly water-soluble substrates. In order to gain 30

insight into the cellular functions involved, we undertook a proteomic analysis of 31

Marinobacter hydrocarbonoclasticus SP17 biofilm developing at the hexadecane-water 32

interface. Biofilm formation on hexadecane led to a global change of the cell physiology 33

involving modulation of the expression of 573 out of 1144 detected proteins when compared 34

with planktonic cells growing on acetate. Biofilm cells overproduced a protein, encoded by 35

MARHY0478 that contains a conserved domain belonging to the family of the outer 36

membrane transporters of hydrophobic compounds. Homologs of MARHY0478 were 37

exclusively found in marine bacteria degrading alkanes or possessing alkane degradation 38

genes and hence presumably constitute a family of alkane transporter specific to marine 39

bacteria. Interestingly, we also found that sessile cells growing on hexadecane overexpressed 40

type VI secretion system components. This secretion system has been identified as a key 41

factor in virulence and in symbiotic interaction with host organisms. This observation is the 42

first experimental evidence of the contribution of a type VI secretion system to environmental 43

adaptation and raises the intriguing question about the role of this secretion machine in alkane 44

assimilation. 45

46

Keywords: Marinobacter hydrocarbonoclasticus SP17; biofilm; alkane degradation; 47

proteomic; type VI secretion system 48

3

49

1. Introduction 50

51

Biofilm formation at the hydrocarbon-water interface has been observed with various 52

alkane degrading strains, e.g. Rhodococcus sp. Q15 [34], Acinetobacter venetianus RAG-1 [2] 53

and Oleiphilus messinensis [9] as well as polycyclic aromatic hydrocarbons (PAHs) 54

degrading strains, e.g. Pseudomonas sp. strain 8909N [19], Sphingomonas sp.CHY-1 [35] and 55

Mycobacterium frederiksbergense LB501T [3]. 56

The ecological significance of these biofims has been demonstrated by a study devoted to the 57

diversity of biofilm communities developing on PAHs [27]. It has been shown that biofilm 58

communities contained a greater diversity of active species and of PAH degradation genes 59

than the planktonic communities enrichment. Furthermore the diversity of active species 60

found in the biofilm closely matched that of the PAH-contaminated soil used as inoculum. 61

These findings suggest that biofilm formation on hydrocarbons is a probable lifestyle in 62

natural ecosystems. 63

Biofilms developing on hydrocarbons present two properties distinguishing them from 64

other biofilms which are their substrate-substratum specificity and their capacity to overcome 65

the low accessibility of hydrophobic substrates. It was observed in the case of PAH-degrading 66

strains that the less soluble the hydrocarbon was, the more cells grew at the PAH-water 67

interface [12,23]. Biofilm formation has been shown to promote growth on hydrocarbons by 68

facilitating interfacial access. Kinetic studies demonstrated that growth at the interface occurs 69

faster than mass transfer rate of hydrocarbons in the absence of bacteria would suggest 70

[4,5,10,13]. Thus, biofilm formation constitutes an efficient adaptive strategy to assimilate 71

hydrocarbon. However, the genetic determinants and the molecular mechanisms underlying 72

the functioning of these biofilms remain poorly understood. In order to identify the proteins 73

4

and cellular functions involved, we undertook a proteomic analysis of the biofilm of 74

Marinobacter hydrocarbonoclasticus SP17 growing at the hexadecane-water interface. This 75

marine alkane degrading bacterium readily forms biofilms on metabolizable hydrophobic 76

organic compounds including n-alkanes from 8 to 28 carbon atoms, whereas in the same 77

condition this strain does not develop biofilms on an inert substratum such as plastic or glass 78

[15]. The doubling time of biofilm cells growing on hexadecane (between 3 and 5 hours 79

depending on the development stage of the biofilm) is similar to that observed on water-80

soluble substrates such as acetate, pyruvate and lactate (P-J Vaysse, unpublished results). 81

Furthermore, the rate of hexadecane degradation dramatically decreases when the biofilm is 82

disorganized by strong shaking, indicating a strong link between biofilm formation and the 83

utilization of alkanes [15]. Theses observations indicate that the biofilm lifestyle must provide 84

M. hydrocarbonoclasticus SP17 with efficient mechanisms to access the hexadecane. In the 85

current study, we undertook a proteomic analysis of a biofilm growing at hexadecane-water 86

interface. The results obtained indicate that adaptation to alkane utilization as carbon and 87

energy source involves a global change in cell physiology. Novel proteins and cellular 88

processes involved in alkane assimilation were revealed. 89

5

2. Material and methods 90

91

2.1 Growth conditions and preparation of protein extracts 92

93

The bacterial strain used in this study was Marinobacter hydrocarbonoclasticus SP17 94

(ATCC 49840). Cultures were carried out in synthetic sea water (SSW) [8]. 95

All chemicals used were from Sigma-Aldrich unless otherwise specified. Biofilm cultures 96

were inoculated with cells growing exponentially on acetate (20 mM) washed twice with one 97

volume of SSW and resuspended to a final optical density at 600nm of 0.1 in 300 ml SSW 98

medium supplemented with 0.2% (v/v) hexadecane in 1 l Erlenmeyer flasks. Biofilms were 99

grown at the hexadecane-SSW interface at 30 °C, under slow shaking (50 rpm). After 35 100

hours incubation, the medium was carefully discarded, the biofilms resuspended in 30 ml of 101

SSW and centrifuged for 20 min at 20,000 g at room temperature. Biofilms were harvested 102

above the supernatant while residual planktonic cells were pelleted. Cells grown on acetate 103

(100 ml) were harvested during exponential growth phase (OD600nm= 0.3) by centrifuging 104

20 min at 20,000 g at room temperature. 105

For protein extraction, cell pellets or biofilms were washed twice with 5 ml of acetone, 106

resuspended in 5 ml of water containing a protease inhibitor cocktail and sonicated on ice for 107

1 min with 500 ms/s pulses at 35 W. Sonicated cell suspensions were then incubated with 40 108

µg of DNase I, 10 µg of RNase A plus 0.01% (v/v) Triton X100 for 30 min at room 109

temperature. Proteins were precipitated for 30 min at 4 °C by adding 500 µl of 100% 110

trichloroacetic acid and then centrifuged 10 min at 20,000g at 4 °C. Protein extracts were 111

washed twice with 1 ml of trichloroacetic acid 10% (v/v) and twice with 1 ml of acetone. 112

Proteins were air dried and dissolved in IEF buffer (urea 8 M, 3-[(3-113

Cholamidopropyl)dimethylammonio]-1-propanesulfonate 4% (w/v), dithiothreitol 60 mM, 114

6

Pharmalyte 3-10 2% (v/v) (Amersham Biosciences) and bromophenol blue 0.0002% (w/v)). 115

Protein concentration was estimated using Biorad Protein assay and finally adjusted to 1 µg/µl 116

with IEF buffer. 117

118

2.2 2D electrophoresis and gel analysis 119

120

Two hundred and thirty micrograms of protein were applied to a 24 cm Immobiline 121

Dry strip with a 3-7 non linear pH gradient (GE Healthcare). Isoelectric focusing and SDS-122

PAGE (12.5%) were carried out using the Multiphor II and Ettan DALTsix systems, 123

respectively (GE Healthcare) according to the manufacturer’s instructions. Gels were stained 124

with Deep Purple Total Protein Stain (GE Healthcare) and then scanned using a Typhoon 125

9200 fluorescent scanner (GE Healthcare). Image analysis, spot detection and matching were 126

carried out using Image Master Platinum software (GE Healthcare) and checked manually. 127

Three replicate gels from three independent experiments were run for each growth condition. 128

The normalized protein amount for each protein spot was defined as the fraction of that spot 129

volume to the total spot’s volume of the gel. Student t-test (P<0.05) and a threshold of 2-fold 130

change were used to determine the proteins significantly differentially expressed between the 131

two conditions. 132

133

2.3 Protein identification by nanoLC-MS/MS 134

135

Protein spots were excised from gels, destained in ammonium bicarbonate 50 mM 136

50% acetonitrile (ACN), rinsed twice in ultrapure water and dehydrated in 100% ACN. After 137

ACN removal by evaporation, gel pieces were dried, rehydrated with a trypsin solution (10 138

ng/ml in 50 mM ammonium bicarbonate) at 4 °C for 10 min and finally incubated overnight 139

7

at 37 °C. The supernatant was collected by two successive extractions with 140

H2O/ACN/HCOOH (47.5/47.5/5), pooled, and concentrated in a vacuum centrifuge to a final 141

volume of 25 µl. The peptide mixture was analyzed by on-line capillary HPLC (LC Packings, 142

Amsterdam, The Netherlands) coupled to a nanospray LCQ Deca XP Ion Trap mass 143

spectrometer (Thermo-Finnigan, San Jose, CA). MS/MS data were acquired using a 2 m/z 144

units ion isolation window and a 35% relative collision energy. Data were searched using 145

SEQUEST through Bioworks 3.3.1 interface (ThermoFinnigan) against the M. 146

hydrocarbonoclasticus SP17 whole genome sequence. Gene numbers, gene function, and 147

functional category are presented according to the unpublished annotated genome. 148

149

2.4 Database searching and sequence analysis 150

151

Sequences similarities were searched against the translation of the non-redundant 152

GenBank CDS database, using the BLAST program [1]. Homologs were defined as proteins 153

giving an alignment with a bits score above 100 (scoring matrix BLOSUM62) and an 154

expected value (E-value) below 1.10-25. Protein domain searches and multiple sequence 155

alignment were carried out using the NCBI-Conserved Domain Database search program with 156

CDD v2.16-27036 PSSMs database (http://www.ncbi.nlm.nih.gov/Structure/cdd/cdd.shtml) 157

[18]. Synteny similarities were searched using the MAGE microbial genome annotation 158

system from the Genoscope Evry France 159

(https://www.genoscope.cns.fr/agc/mage/wwwpkgdb/MageHome). 160

8

161

3. Results and discussion 162

163

3.1. Comparison of protein patterns of biofilm cells growing on hexadecane with planktonic 164

cells growing on acetate 165

166

In order to characterize the molecular mechanisms involved in the development of 167

biofilm on alkanes, we compared the proteomes of biofilm growing on hexadecane and 168

planktonic cells growing on acetate, hereafter referred to as BH and PA respectively (Fig. 1). 169

A total of 1144 spots appearing on all three replicates of at least one growth condition were 170

detected. 576 spots (50%) were found to change significantly between PA and BH. Among 171

these proteins, 245 were overexpressed in BH and 331 in PA. 81% of the differentially 172

expressed protein had an induction ratio above 10. This high proportion of differentially 173

expressed proteins is unusual when compared to other similar proteomic analyses. For 174

instance, a study of Alcanivorax burkomensis SK2 comparing growth on hexadecane with 175

growth on pyruvate revealed 97 proteins differentially expressed [24]. However, A. 176

burkomensis SK2 protein extracts were prepared from whole cultures consisting of a mixture 177

of biofilm cells and detached planktonic cells. Therefore, the expression protein profiles 178

might reflect the average protein expression levels from different cellular states, thus 179

lessening the actual expression fold-change of one specific condition. Furthermore, in our 180

study, proteins were revealed using a fluorescent dye with a higher sensitivity and larger 181

dynamic range than Coomassie blue. 182

The high number of differentially expressed proteins between the biofilm cells 183

growing on hexadecane and those growing on acetate signify that growth on hexadecane 184

involve a global change in the cell physiology, requiring numerous cellular functions. This 185

9

could be explained by the change of two major factors between the compared conditions: the 186

lifestyle (biofilm or planktonic) and the carbon source (hexadecane or acetate). In many 187

species, biofilm formation has been shown to require a large number of protein functions. The 188

biofilm lifestyle induces changes in the environmental conditions encountered by cells, such 189

as the formation of nutrient microgradients. The adaptation to these new conditions requires 190

most likely a great number of proteins [28]. On the other hand, the response to alkanes 191

involves the modulation of the expression of numerous proteins required for the transport, the 192

metabolism or to avoid toxic effects of these compounds [31]. A total of 58 spots, that could 193

unambiguously be assigned to a single protein, were kept for further analysis (Table 1). The 194

major type IV pilus subunit, PilA (encoded by MARHY2564) was under-produced in BH. 195

Type IV pili have been shown to be involved in host cell adhesion, biofilm formation, DNA 196

uptake and twitching motility [21]. A DNA microarray analysis showed that pilin genes were 197

repressed in the biofilm of Pseudomonas aeruginosa [33]. Furthermore, the type IV pilus of 198

P. aeruginosa has been shown to have multiple effects on biofilm formation mainly through 199

twitching motility and adhesion [11,20]. 200

Proteins involved in nutrient transport across cellular envelopes constitute a large class 201

of proteins overproduced in BH. This includes components of phosphate (encoded by 202

MARHY3535 pstS) and thiosulfate (encoded by MARHY2019 cysP) ABC transporters. Iron 203

uptake and transport proteins were represented by FhuE, FbpA and CirA encoded by 204

MARHY1035, MARHY2192 and MARHY3135 respectively. Four other proteins coding for 205

putative ABC transporters and porins (encoded by MARHY0256, MARHY0299, 206

MARHY3277, MARHY34332 and MARHY3166) were also found, although their actual 207

function remain uncertain. Increased capacities in solute transport may reflect an adaptation to 208

the constraints imposed by the biofilm lifestyle. Indeed, the increase in biofilm thickness is 209

thought to hinder nutrient penetration into the deepest layers of the biofilm. Thus, cells would 210

10

require an increased capacity in nutrient uptake and transport. This is particularly important in 211

seawater where iron and phosphate are in limiting concentrations. Micronutrients such as iron 212

and inorganic phosphate are also known to strongly influence biofilm development by acting 213

as environmental cues regulating biofilm formation [25]. 214

215

3.2. Redirection of carbon flux in biofilm cells growing on hexadecane 216

Most of the enzymes corresponding to the CO2-releasing steps of the tricarboxylic acid 217

cycle (TCA cycle) (encoded by MARHY0078 idh, MARHY2120 sucD, MARHY2121 sucC, 218

MARHY2126 sdhA) were found to be down regulated in BH. In addition, the gene encoding 219

malate synthase (MARHY1458 glcB), catalyzing the conversion of glyoxylate into malate, 220

was upregulated. This modulation of TCA cycle enzymes suggests a stimulation of the 221

glyoxylate bypass in BH. This anaplerotic pathway allows the replenishment of cells in the 222

metabolic intermediates necessary for the synthesis of their cellular components when acetyl-223

CoA is the only carbon source available in the cell [16]. The activation of the glyoxylate 224

bypass under a hexadecane diet compared to a diet of pyruvate or glucose has been already 225

observed in proteomic analyses on A. burkomensis SK2 and Geobacillus thermodenitrificans 226

[6,24]. In our study, stimulation of the glyoxylate bypass was observed under hexadecane 227

conditions compared with acetate. In hexadecane and acetate conditions, the intracellular 228

carbon source is acetyl-CoA hence the glyoxylate pathway is required in both conditions. The 229

explanation of the glyoxylate bypass stimulation could lie in the fact that hexadecane is a 230

more energetic substrate than acetate. In fact the breakdown of hexadecane to acetyl-CoA 231

produces energy whereas the conversion of acetate to acetyl-CoA does not. Augmenting the 232

flux of carbon through the glyoxylate bypass on hexadecane could allow cells to restore the 233

balance between carbon assimilation and energy production. However interpretation of 234

11

proteomic data with regard to metabolic flux should be taken with caution, since metabolic 235

pathways can be regulated at the level of enzyme activity, in addition to gene expression. 236

Fatty acid biosynthesis genes fabA, fabB and fabF (MARHY3086, MARHY3087 and 237

MARHY1438, respectively) were found to be down regulated in BH. This corroborates the 238

fact that in cells growing on hexadecane the main fatty acids of cellular lipids were derived 239

from the oxidation of alkanes [26]. 240

241

3.3. A type VI secretion system is overproduced by biofilm cells growing on hexadecane 242

243

Three proteins sharing similarities with type VI secretion system (T6SS) subunits, 244

encoded by MARHY3623, MARHY3634 and MARHY3635, were overexpressed in BH. 245

MARHY3634 and MARHY3635 were among the most abundant protein detected in BH, with 246

0.2 and 0.7 % of total protein respectively. MARHY3634, MARHY3635 and MARHY3623 247

are localized within a cluster of 16 genes (from MARHY3635 to MARHY3620) that are 248

transcribed in the same direction. All members of the M. hydrocarbonoclasticus SP17 cluster 249

share similarity with genes found in T6SS clusters from other bacteria including Vibrio 250

cholerae, P. aeruginosa and Escherichia coli. The best synteny conservation with a 251

characterized T6SS was observed with Vibrio cholerae N16961 T6SS gene cluster (Fig. 2). 252

Homologs of component of T6SS already characterized, i.e. Dot, IcmF, and ClpV (a 253

subfamily of ClpB ATPase), were also found in M. hydrocarbonoclasticus SP17 cluster 254

(MARHY3626, MARHY3621 and MARHY3625). Vgr and Hcp are two proteins secreted 255

through T6SS. One homolog of Vgr is found (MARH3620) within this cluster while other 256

homologs of Vgr and Hcp are found elsewhere on the M. hydrocarbonoclasticus SP17 257

chromosome [7] (Fig. 3). Experimental evidence of the functionality of homologs of 258

MARHY3635 and 3634 was provided by studies of the pathogen Edwardsielle tarda [22]. 259

12

Deletion mutants of evpA and evpB (MARHY3635 and MARHY3634, respectively) were 260

impaired in protein secretion and showed reduced virulence in blue gourami fish. Based on 261

the high degree of sequence similarity and synteny conservation between characterized T6SS 262

clusters, we propose that the genes from MARHY3635 to MARHY3620 constitute a T6SS 263

gene cluster. The functionality of this T6SS is supported by the detection of three proteins 264

encoded by this cluster. 265

Although type VI secretion systems have been identified as key factors in virulence of 266

pathogenic bacteria and in symbiotic interaction with host organisms, in silico analyses have 267

revealed their presence in many species that are not considered as pathogens or symbionts. 268

This led to the hypothesis that T6SS may also contribute to environmental adaptation [29]. 269

Our data provide the first experimental evidence of the production of a T6SS during biofilm 270

development by an environmental bacterium. The role of T6SS in biofilm formation during 271

growth on hexadecane by M. hydrocarbonoclasticus SP17 remains puzzling. T6SS may be 272

required to secrete proteins needed for biofilm formation or for efficient assimilation of 273

carbon source difficult to access. 274

275

3.4. Evidence for an alkane transporter family specific to marine bacteria 276

277

Two proteins produced in large quantities in BH, encoded by MARHY0478 and 278

MARHY0477, drew our attention because their sequences were found to be conserved 279

exclusively in marine species capable to use alkanes as a carbon source or possessing at least 280

one copy of an alkane hydroxylase gene. The phylogenetic distribution of MARHY0478 and 281

MARHY0477 homologs is rather restricted since they are found only in two orders of the 282

Gammaproteobacteria: the Oceanospirillales and the Alteromonadales. The degree of peptide 283

sequence identity varies between 29% and 98%. Interestingly, these proteins are found in 284

13

Alcanivorax borkumensis SK2, a key alkane degrader in polluted seawater. Several strains 285

possess multiple homologs of MARHY0478 and MARHY0477 with a maximum of six in 286

Alcanivorax sp DG881 (Table 2). 287

In M. hydrocarbonoclasticus SP17, MARHY0478 and MARHY0477 are separated by only 288

11 base pairs and are transcribed in the same direction. This spatial organization is indicative 289

of a putative operon. Such an operon would suggest that the products of these genes are 290

involved in the same function. An operon structure is likely to be conserved in M. 291

hydrocarbonoclasticus VT8 and A. burkumensis SK2 since these strains contain homologs of 292

MARHY0477 and MARHY0478 that are adjacent and transcribed in the same direction. 293

The strong expression of MARHY0478 and MARHY0477 in biofilm on hexadecane indicates 294

that they might play an important role in alkane assimilation. In order to gain more insight 295

into their function, conserved domains in the two proteins were searched against the NCBI 296

Conserved Domain Database using RPS-BLAST [18]. Only the protein encoded by 297

MARHY0478 gave a positive hit (E-value = 10-13) with the pfam03349 domain family. This 298

conserved domain is found in outer membrane proteins transporting hydrophobic compounds 299

out of Gram negative bacteria. This family includes the monoaromatic hydrocarbon transport 300

proteins TodX from Pseudomonas putida F1 and TbuX from Ralstonia pickettii PKO1, and 301

the long chain fatty acid tranporter FadL from Escherichia coli [14,30,32]. Fig. 3 shows the 302

multiple peptide sequence alignment for the conserved domain including the amino acid 303

sequence deduced from MARHY0478, TodX, TbuX and FadL. This alignment is based upon 304

the frequency of each amino acid substitution in the set of protein used to create the conserved 305

domain model [17]. Peptide encoded by MARHY0478 exhibited a clear sequence similarity 306

with the four hydrophobic compounds transporters. In view of this significant domain 307

conservation, it is very likely that MARHY0478 encodes for a hydrophobic compound 308

transporter. Given that MARHY0478 was induced during growth on hexadecane, and that 309

14

homologs are only found in marine alkane degrading strains or putative alkane degrading 310

strains, we hypothesized that MARHY0478 and its homologs would constitute a family of 311

alkane transporter specific to marine bacteria. The production in large quantity of an alkane 312

transporter in biofilm cells would contribute to the high growth rate observed on alkane 313

despite its low solubility. 314

Polypeptides encoded by MARHY0477 and their homologs define a protein family 315

whose function is likely to be related to MARHY0478, as members of the two families co-316

occur in the same strains and occasionally form a putative operon. The ecological significance 317

of the MARHY0478 and MARHY0477 families certainly deserve closer inspection as they 318

are present in genera known to be the main actors in hydrocarbon degradation in marine 319

environments. 320

321

15

322

Acknowledgments 323

We thank Dr Rizard Lobinsky, Dr Hugues Prudhomme and the Pôle Protéomique of 324

the Bordeaux 2 University for protein identification. We also thank Dr Pierre Sivadon and Dr 325

Anne Fahy for useful discussions and critical reading of the manuscript. We gratefully 326

acknowledge the 6th European Framework Programme, Contract 018391 FACEIT, the 327

National Program ANR “ECCO” INDHYC project, the CNRS program Ingénierie 328

Ecologique, the région Aquitaine and the département des Pyrénées Atlantiques for financial 329

support. 330

16

References 331

[1] Altschul, S.F., Madden, T.L., Schaffer, A.A., Zhang, J., Zhang, Z., Miller, W., Lipman, 332

D.J. (1997) Gapped BLAST and PSI-BLAST: a new generation of protein database search 333

programs. Nucleic Acids Res. 25, 3389-3402. 334

[2] Baldi, F., Ivosevic, N., Minacci, A., Pepi, M., Fani, R., Svetlicic, V., Utlic, V. (1999) 335

Adhesion of Acinetobacter venetianus to diesel fuel droplets studied with in situ 336

electrochemical and molecular probes. Appl. Environ. Microbiol. 65, 2041-2048. 337

[3] Bastiaens, L., Springael, D., Wattiau, P., Harms, H., DeWachter, R., Verachtert, H., Diels, 338

L. (2000) Isolation of adherent polycyclic aromatic hydrocarbon (PAH)-degrading bacteria 339

using PAH-sorbing carriers. Appl. Environ. Microbiol. 66, 1834-1843. 340

[4] Bouchez-Naitali, M., Blanchet, D., Bardin, V., Vandecasteele, J.-P. (2001) Evidence for 341

interfacial uptake in hexadecane degradation by Rhodococcus equi: The importance of cell 342

flocculation. Microbiology 147, 2537-2543. 343

[5] Calvillo, Y.M., Alexander, M. (1996) Mechanism of microbial utilization of biphenyl 344

sorbed to polyacrylic beads. Appl. Microbiol. Biotechnol. 45, 383-390. 345

[6] Feng, L., Wang, W., Cheng, J., Ren, Y., Zhao, G., Gao, C., Tang, Y., Liu, X., et al. (2007) 346

Genome and proteome of long-chain alkane degrading Geobacillus thermodenitrificans 347

NG80-2 isolated from a deep-subsurface oil reservoir. Proc. Natl. Acad. Sci. U S A 104, 348

5602-5607. 349

[7] Filloux, A., Hachani, A., Bleves, S. (2008) The bacterial type VI secretion machine: Yet 350

another player for protein transport across membranes. Microbiology 154, 1570-1583. 351

[8] Gauthier, M.J., Lafay, B., Christen, R., Fernandez, L., Acquaviva, M., Bonin, P., Bertrand, 352

J.-C. (1992) Marinobacter hydrocarbonoclasticus gen. nov., sp. nov., a new, extremely 353

halotolerant, hydrocarbon-degrading marine bacterium. Int. J. Syst. Bacteriol. 42, 568-576. 354

17

[9] Golyshin, P.N., Chernikova, T.N., Abraham, W.-R., Lünsdorf, H., Timmis, K.N., 355

Yakimov, M.M. (2002) Oleiphilaceae fam. nov., to include Oleiphilus messinensis gen. nov., 356

sp. nov., a novel marine bacterium that obligately utilizes hydrocarbons. Int. J. Syst. Evol. 357

Microbiol. 52, 901-911. 358

[10] Harms, H., Zehnder, A.J.B. (1995) Bioavailability of sorbed 3-chlorodibenzofuran. Appl. 359

Environ. Microbiol. 61, 27-33. 360

[11] Heydorn, A., Ersbøll, B., Kato, J., Hentzer, M., Parsek, M.R., Tolker-Nielsen, T., 361

Givskov, M., Molin, S. (2002) Statistical analysis of Pseudomonas aeruginosa biofilm 362

development: Impact of mutations in genes involved in twitching motility, cell-to-cell 363

signaling, and stationary-phase sigma factor expression. Appl. Environ. Microbiol. 68, 2008-364

2017. 365

[12] Johnsen, A.R., Karlson, U. (2004) Evaluation of bacterial strategies to promote the 366

bioavailability of polycyclic aromatic hydrocarbons. Appl. Microbiol. Biotechnol. 63, 452-367

459. 368

[13] Johnsen, A.R., Wick, L.Y., Harms, H. (2005) Principles of microbial PAH-degradation 369

in soil. Environ. Pollut. 133, 71-84. 370

[14] Kahng, H.Y., Byrne, A.M., Olsen, R.H., Kukor, J.J. (2000) Characterization and role of 371

tbuX in utilization of toluene by Ralstonia pickettii PKO1. J. Bacteriol. 182, 1232-1242. 372

[15] Klein, B., Grossi, V., Bouriat, P., Goulas, P., Grimaud, R. (2008) Cytoplasmic wax ester 373

accumulation during biofilm-driven substrate assimilation at the alkane-water interface by 374

Marinobacter hydrocarbonoclasticus SP17. Res. Microbiol. 159, 137-144. 375

[16] Kornberg, H.L. (1966) The role and control of the glyoxylate cycle in Escherichia coli. 376

Biochem. J. 99, 1-11. 377

18

[17] Marchler-Bauer, A., Anderson, J.B., Derbyshire, M.K., DeWeese-Scott, C., Gonzales, 378

N.R., Gwadz, M., Hao, L., He, S., et al. (2007) CDD: a conserved domain database for 379

interactive domain family analysis. Nucleic Acids Res. 35, D237-D240. 380

[18] Marchler-Bauer, A., Anderson, J.B., Chitsaz, F., Derbyshire, M.K., Deweese-Scott, C., 381

Fong, J.H., Geer, L.Y., Geer, R.C., et al. (2009) CDD: Specific functional annotation with the 382

Conserved Domain Database. Nucleic Acids Res. 37, D205-D210. 383

[19] Mulder, H., Breure, A.M., Van Honschooten, D., Grotenhuis, J.T.C., Van Andel, J.G., 384

Rulkens, W.H. (1998) Effect of biofilm formation by Pseudomonas 8909N on the 385

bioavailability of solid naphthalene. Appl. Microbiol. Biotechnol. 50, 277-283. 386

[20] O'Toole, G.A., Kolter, R. (1998) Flagellar and twitching motility are necessary for 387

Pseudomonas aeruginosa biofilm development. Mol. Microbiol. 30, 295-304. 388

[21] Proft, T., Baker, E.N. (2009) Pili in Gram-negative and Gram-positive bacteria - 389

structure, assembly and their role in disease. Cell. Mol. Life Sci. 66, 613-635. 390

[22] Rao, P.S., Yamada, Y., Tan, Y.P., Leung, K.Y. (2004) Use of proteomics to identify 391

novel virulence determinants that are required for Edwardsiella tarda pathogenesis. Mol. 392

Microbiol. 53, 573-586. 393

[23] Rodrigues, A.C., Wuertz, S., Brito, A.G., Melo, L.F. (2005) Fluorene and phenanthrene 394

uptake by Pseudomonas putida ATCC 17514: Kinetics and physiological aspects. Biotechnol. 395

Bioeng. 90, 281-289. 396

[24] Sabirova, J.S., Ferrer, M., Regenhardt, D., Timmis, K.N., Golyshin, P.N. (2006) 397

Proteomic insights into metabolic adaptations in Alcanivorax borkumensis induced by alkane 398

utilization. J. Bacteriol. 188, 3763-3773. 399

[25] Singh, P.K., Parsek, M.R., Greenberg, E.P., Welsh, M.J. (2002) A component of innate 400

immunity prevents bacterial biofilm development. Nature 417, 552-555. 401

19

[26] Soltani, M., Metzger, P., Largeau, C. (2004) Effects of hydrocarbon structure on fatty 402

acid, fatty alcohol, and beta-hydroxy acid composition in the hydrocarbon-degrading 403

bacterium Marinobacter hydrocarbonoclasticus. Lipids 39, 491-505. 404

[27] Stach, J.E., Burns, R.G. (2002) Enrichment versus biofilm culture: a functional and 405

phylogenetic comparison of polycyclic aromatic hydrocarbon-degrading microbial 406

communities. Environ. Microbiol. 4, 169-182. 407

[28] Stewart, P.S., Franklin, M.J. (2008) Physiological heterogeneity in biofilms. Nat. Rev. 408

Microbiol. 6, 199-210. 409

[29] Tseng, T.-T., Tyler, B.M., Setubal, J.C. (2009) Protein secretion systems in bacterial-host 410

associations, and their description in the Gene Ontology. BMC Microbiology 9 (Suppl 1):S2, 411

1-9. 412

[30] van den Berg, B., Black, P.N., Clemons, W.M., Jr., Rapoport, T.A. (2004) Crystal 413

structure of the long-chain fatty acid transporter FadL. Science 304, 1506-1509. 414

[31] Van Hamme, J.D., Singh, A., Ward, O.P. (2003) Recent advances in petroleum 415

microbiology. Microbiol. Mol. Biol. Rev. 67, 503-549. 416

[32] Wang, Y., Rawlings, M., Gibson, D.T., Labbe, D., Bergeron, H., Brousseau, R., Lau, 417

P.C. (1995) Identification of a membrane protein and a truncated LysR-type regulator 418

associated with the toluene degradation pathway in Pseudomonas putida F1. Mol. Gen. 419

Genet. 246, 570-579. 420

[33] Whiteley, M., Bangera, M.G., Bumgarner, R.E., Parsek, M.R., Teitzel, G.M., Lory, S., 421

Greenberg, E.P. (2001) Gene expression in Pseudomonas aeruginosa biofilms. Nature 413, 422

860-864. 423

[34] Whyte, L.G., Slagman, S.J., Pietrantonio, F., Bourbonniere, L., Koval, S.F., Lawrence, 424

J.R., Inniss, W.E., Greer, C.W. (1999) Physiological adaptations involved in alkane 425

20

assimilation at a low temperature by Rhodococcus sp. strain Q15. Appl. Environ. Microbiol. 426

65, 2961-2968. 427

[35] Willison, J.C. (2004) Isolation and characterization of a novel sphingomonad capable of 428

growth with chrysene as sole carbon and energy source. FEMS Microbiol. Lett. 241, 143-150. 429

430

21

Legends to figures 431

432

Fig. 1. Two-dimensional gel electrophoresis of protein extracts of M. hydrocarbonoclasticus 433

SP17. Planktonic cells grown on acetate (top image) and biofilm cells grown on hexadecane 434

(bottom image). Spots of identified proteins are numbered. Molecular weights (in kDa) are 435

indicated on the left side. 436

437

Fig. 2. Organization of the genes encoding for type VI secretion system of M. 438

hydrocarbonoclasticus SP17 (top) and Vibrio cholerae N16961 (bottom). The values between 439

homologous genes refer to the percent peptide sequence identity of their products. Accession 440

number : MARHY3635 to MARHY3620 (FP475883 to FP47598), MARHY0176 441

(FP475953), MARHY1076 (FP475938) and MARHY2494 (FP475918) 442

443

Fig. 3. Multiple sequence alignment of FadL, TbuX, TodX and translated MARHY0478. 444

Sequence alignment was performed using the web-based tool CD search using the PSSM 445

112176 scoring matrix. Conserved residues are shown in grey boxes and identical residues in 446

black boxes. Secondary structures elements of FadL were retrieved from the PDB database 447

(PDB ID:1t16) and are shown above the sequences. 448

449

22

Tables Table 1

Identification of proteins differentially expressed in biofilm on hexadecane compared to planktonic cells on acetate.

Spot NO.

CDS Gene Product Differential abundancea

Amount in planktonic/acetate

conditionb

Amount in biofilm/hexadecane

conditionb

Gene accession numberc

Transport and binding protein

1 MARHY1035 fhuE Outer membrane receptor for ferric iron uptake +5.01 0.0102 0.0513 FP475940 2 MARHY2192 fbpA Iron(III) ABC transporter, periplasmic iron(III)-binding

protein +3.89 0.1233 0.4793 FP475923

3 MARHY3135 cirA Ferric iron-catecholate outer membrane transporter BH ND 0.079 FP475909 4 MARHY2019 cysP Thiosulfate transporter subunit +3.01 0.1170 0.3525 FP475928 5 MARHY3535 pstS ABC phosphate transporter, periplasmic component +10.74 0.1018 1.0934 FP475901 6 MARHY3277 _ ABC-type metal ion transporter, periplasmic component BH ND 0.0831 FP475905 7 MARHY1478 _ ABC-type branched-chain amino acid transporter, periplasmic

component -4.32 0.0882 0.0204 FP475934

Cellular processes

8 MARHY2564 pilA Fimbrial protein precursor -2.44 0.6983 0.2861 FP475917 9 MARHY2994 tpm Thiopurine methyltransferase PA 0.0551 ND FP475913

Information transfer 10 MARHY3200 greA Transcription elongation factor BH ND 0.0843 FP475907

Protein fate 11 MARHY0922 _ FKBP-type peptidyl-prolyl cis-trans isomerase PA 0.0418 ND FP475944 12 MARHY0942 dsbC Protein disulfide isomerase II +4.44 0.0363 0.1611 FP475943 13 MARHY0958 mucD Serine protease +4.05 0.0517 0.2091 FP475942 14 MARHY1798 slyD FKBP-type peptidyl-prolyl cis-trans isomerase PA 0.0584 ND FP475931

23

Spot NO.

CDS Gene Product Differential abundancea

Amount in planktonic/acetate

conditionb

Amount in biofilm/hexadecane

conditionb

Gene accession numberc

Stress response

15 MARHY3764 ahpF Alkyl hydroperoxide reductase f-subunit BH ND 0.0369 FP475881

Cofactor biosynthesis 16 MARHY2568 coaE Dephospho-CoA kinase PA 0.0533 ND FP475916

Aminoacid biosynthesis 17 MARHY1723 asd Aspartate-semialdehyde dehydrogenase PA 0.0655 ND FP475932 18 MARHY3112 hisH Glutamine amidotransferase BH ND 0.1733 FP475910 19 MARHY0406 lysA Diaminopimelate decarboxylase PA 0.0901 ND FP475949 20 MARHY2259 serC 3-phosphoserine/phosphohydroxythreonine aminotransferase BH ND 0.1327 FP475920

Nucleotide biosynthesis 21 MARHY2255 cmk Cytidylate kinase BH ND 0.0691 FP475921 22 MARHY2158 ndk Multifunctional nucleoside diphosphate kinase and

apyrimidinic endonuclease and 3'-phosphodiesterase BH ND 0.0836 FP475924

23 MARHY3664 pyrC Aspartate carbamoyltransferase, non-catalytic chain BH ND 0.0533 FP475882 24 MARHY2249 pyrF Orotidine-5'-phosphate decarboxylase +3.35 0.0582 0.1948 FP475922

Central intermediary metabolism 25 MARHY3262 exaC NAD+ dependent acetaldehyde dehydrogenase -3.44 0.0968 0.0281 FP475906

Energy metabolism 26 MARHY1444 acnB Bifunctional aconitate hydratase 2 and 2-methylisocitrate

dehydratase -2.89 0.9104 0.3147 FP475936

27 MARHY0078 idh Isocitrate dehydrogenase -3.13 0.2014 0.0643 FP475954 28 MARHY2120 sucD Succinyl-CoA synthetase, NAD(P)-binding, alpha-subunit -3.2 0.8657 0.2708 FP475927 29 MARHY2121 sucC Succinyl-CoA synthetase, beta-subunit -6.97 0.1894 0.0272 FP475926 30 MARHY2126 sdhA Succinate dehydrogenase, flavoprotein subunit -4.35 0.3639 0.0837 FP475925 31 MARHY0774 petA Ubiquinol-cytochrome c reductase iron-sulfur subunit +6 0.0449 0.2696 FP475945 32 MARHY1458 glcB Malate synthase G BH ND 0.0203 FP475935 33 MARHY1802 etfA Electron transfer flavoprotein subunit, FAD-binding +3.13 0.1707 0.5347 FP475930 34 MARHY1815 paaG Enoyl-CoA hydratase PA 0.073 ND FP475929

24

Spot NO.

CDS Gene Product Differential abundancea

Amount in planktonic/acetate

conditionb

Amount in biofilm/hexadecane

conditionb

Gene accession number

Fatty acid biosynthesis

35 MARHY1009 acs Acetyl-CoA synthetase PA 0.0435 ND FP475941 36 MARHY3086 fabA Beta-hydroxydecanoyl thioester dehydrase PA 0.1247 ND 37 MARHY3087 fabB 3-oxoacyl-[acyl-carrier-protein] synthase I -2.54 0.5215 0.2056 FP475911 38 MARHY1438 fabF 3-oxoacyl-[acyl-carrier-protein] synthase II -4.04 0.1566 0.0388 FP475937 39 MARHY1579 prpE Propionate-CoA ligase -5.75 0.1203 0.0209 FP475933

Uncertain or unknown function 40 MARHY0256 _ Probable ABC transporter, ATPase subunit PA 0.2227 ND FP475952 41 MARHY0299 _ PropableTRAP dicarboxylate transporter BH ND 0.123 FP475951 42 MARHY3613 _ Probable PspA protein +3.5 0.1716 0.6 FP475899 43 MARHY3634 _ Probable type VI secretion system subunit BH ND 0.2073 FP475884 44 MARHY3635 _ Probable type VI secretion system subunit +6.68 0.1118 0.7467 FP475883 45 MARHY3432 _ Probable TonB-dependent receptor +10.99 0.0193 0.2118 FP475902 46 MARHY3289 _ Probable porin PA 0.4185 ND FP475904 47 MARHY0478 _ Probable hydrophobic compounds transporter +11.23 0.0369 0.4144 FP475946 48 MARHY0477 _ Conserved protein +5.62 0.0369 0.2076 FP475947 49 MARHY0460 _ Conserved protein BH ND 0.696 FP475948 50 MARHY1073 _ Conserved protein PA 0.0385 ND FP475939 51 MARHY2963 _ Conserved protein +9.21 0.0326 0.3001 FP475914 52 MARHY3166 _ Conserved protein -14.81 3.612 0.2445 FP475908 53 MARHY3295 _ Conserved protein BH ND 0.0961 FP475903 54 MARHY0333 _ Conserved protein BH ND 0.1249 FP475950 55 MARHY3623 _ Probable type VI secretion system subunit +4.71 0.0142 0.0671 FP475895 56 MARHY3550 _ Conserved protein PA 0.1351 ND FP475900 57 MARHY2686 _ Conserved protein +22.01 0.0924 2.0331 FP475915 58 MARHY2432 _ Conserved protein PA 0.0642 ND FP475919

aPositive values represent overexpression in biofilm/hexadecane condition. Negative values represent underexpression in biofilm/hexadecane

condition. BH means that the protein is solely detected in biofilm/hexadecane condition. PA means that the protein is solely detected in

planktonic/acetate condition. bAverage percent of spot volume relative to total spot volume. ND, not detected.

25

Table 2

Phylogenetic distribution of MARHY0477 and MARHY0478 homologs.

Sequences similar to translated MARHY0477 and MARHY0478 were searched against the non-redundant GenBank CDS translations database,

using the BLAST program. aND not determined. bNumber of homologous genes and their percent peptide sequence identity in brackets. calk and

P450 indicate the presence of alkane hydroxylase and cytochrome P450 alkane monooxygenase genes respectively.

Strain Provenance Alkane

degradationa MARHY0477

homologsb MARHY0478

homologsb Alkane degradation

genesc

Marinobacter hydrocarbonoclasticus SP17 Marine sediment, Mediterranean Sea yes - - 2 alk, 2 P450

Marinobacter hydrocarbonoclasticus VT8 Oil well off the Vietnamese coast yes 2 (74%, 33%) 3 (98%, 74%, 74%) 3 alk, 1 P450

Marinobacter algicola DG893 Culture of a dinoflagellate yes 1 (59%) 1 (83%) 2 alk

Alcanivorax borkumensis SK2 Sea water, North Sea yes 2 (34%, 32%) 1 (40%) 2 alk, 1 P450

Alcanivorax sp. DG881 Culture of marine algae yes 6 (35%, 32%, 33%, 31%, 31%, 29%)

6 (54%, 54%, 50%, 48%, 46%, 42%)

2 alk, 2 P450

Bermanella marisrubri Sea water, Red Sea ND 1 (31%) 1 (45%) 1 alk

Moritella sp. PE36 Deep sea floor, Pacific Ocean ND 1 (29%) 1( 39%) 1 alk

26

Figures Fig. 1

27

Fig. 2

28

Fig. 3